TDP-43 Role in Chromatin Remodeling and Transcription: Comparison
Please note this is a comparison between Version 2 by Camila Xu and Version 1 by Juliette Gimenez.

TDP-43 gained momentum in the neurodegeneration field when it was first discovered that almost all amyotrophic lateral sclerosis (ALS) cases and as many as half of frontotemporal dementia (FTD) cases present pathological ubiquitinated inclusions of TDP-43. Its involvement in chromatin silencing and nuclear/cytoplasmic shuttling constitute convergent key findings from several biological screens, and several crucial epigenetic factors appear to be able to modify TDP-43-induced degeneration. TDP-43 activity at the chromatin level and its implication in the regulation of DNA transcription and stability -such as DNA repair and regulation of retrotransposons activity- are further supported by a continuously growing amount of studies.

  • TDP-43
  • TARDBP
  • neurodegeneration
  • ALS
  • FTD/FTLD
  • epigenetics

1. Introduction

Epigenetic mechanisms in both general and specific neurodegenerative diseases are gaining momentum [1,2,3,4,5,6][1][2][3][4][5][6].
TDP-43 gained momentum in the neurodegeneration field when it was first discovered that almost all amyotrophic lateral sclerosis (ALS) cases and as many as half of frontotemporal dementia (FTD) cases present pathological ubiquitinated inclusions of TDP-43 [7,8][7][8]. Since then, deregulated TDP-43 has been described in several neurodegenerative diseases with different degrees of penetration, from ALS (97%) to FTD (45–50%) to Alzheimer’s disease (AD, 40–50%) (as reviewed in [9]). TDP-43 aggregates have also been found in patients with Huntington’s disease [10[10][11],11], in the brains of humans following traumatic brain injury (TBI) [12,13][12][13], and in an increasing list of neurodegenerative or aging-related diseases [14]. More recently, alterations in TDP-43 regulation/aggregation have also been described in many patients affected by inclusion body myositis [15] and in models of Niemann–Pick disease [16[16][17],17], thus extending the list of TDP-43-associated diseases beyond the strict neurodegenerative spectrum.
At the functional level, TDP-43 is a highly conserved protein involved in the regulation of RNA processing. Mechanistically, TDP-43 is able to bind with different affinity both single-stranded (ss)RNA and ssDNA but also double-stranded DNA (dsDNA). From a structural point of view, the binding of TDP-43 to nucleic acids is mediated by two RNA recognition motif (RRM) domains in its N-terminal region [15,16,17,18,19,20,21,22,23][15][16][17][18][19][20][21][22][23].
In agreement with its strong affinity for nucleic acids, TDP-43 localization is mainly nuclear, and cytoplasmic aggregation is attributed to pathological processes. However, it is now well recognized that TDP-43 is capable of shuttling back and forth from the nucleus to the cytoplasm, even under normal physiological conditions [24]. On the one hand, in neurons, TDP-43 is able to shift to the cytoplasm and travel along the axons to bring mRNA to the synapses for local translation, a function notably impaired in stem-cell-derived motor neurons from ALS patients bearing TDP-43 ALS-causing mutations [25,26,27,28][25][26][27][28]. On the other hand, pathological TDP-43 accumulates in dense cytoplasmic inclusions that include full-length protein and protease cleavage products such as C-terminal TDP-43 fragments (CTFs), as well as abnormally phosphorylated and ubiquitinated proteins [9,29,30,31][9][29][30][31]. When irreversibly aggregated in the cytoplasm, it is believed that the protein is unable to perform its normal functions and thus leads to a loss-of-function scenario (although gain-of-function consequences may be present).
Being a DNA/RNA-binding protein, TDP-43 is highly involved in many aspects of RNA metabolism, such as the control of alternative splicing (AS), microRNA (miRNA) processing, and messenger RNA (mRNA) stability and transport [32,33,34,35,36][32][33][34][35][36]. Nonetheless, recent reports have broadened its function to the regulation of a wide range of chromatin features, from gene transcriptional regulation to DNA repair, passing by chromatin shaping, and the control of retrotransposons for DNA stability. However, because of the abundance of biological functions, it is still yet unclear which ones are early/central to neurodegenerative processes.

2. TDP-43 Role in Chromatin Remodeling and Transcription

During the last decade, TDP-43 has been mostly studied for the functions linked to its RNA-binding properties. Notwithstanding this focus, initial studies established the capacity of TDP-43 to bind to ssDNA TG repeats with at least the same efficiency as it does to UG-repeated sequences within RNA [15,16[15][16][19][22],19,22], while binding to other motifs, e.g., to the ssDNA HIV TAR motif from which its name is derived, performed at a slower association rate and at an even slower dissociation rate than it does to (TG)6 stretch [19,20][19][20]. Instead, its ability to bind dsDNA has been documented in different studies [15,20,21][15][20][21] and notably regarding free dsDNA ends [23]. More recently, additional studies have found TDP-43 to be able to specifically bind DNA sequences in promoter regions and affect the expression of several genes [15,19,37,38,39[15][19][37][38][39][40][41][42][43][44][45][46][47][48][49],40,41,42,43,44,45,46,47,48,49], discussed in detail below. Although the function of TDP-43 on chromatin is yet to be fully understood, it is now very clear that the toxic effects of altered TDP-43 can affect chromatin homeostasis.

2.1. TDP-43 Is a Global Chromatin Modifier

As already mentioned, the function of TDP-43 as a more general transcriptional activator/repressor has been known for about a decade [40]. Its involvement in chromatin silencing and nuclear/cytoplasmic shuttling constitute convergent key findings from several biological screens, and several crucial epigenetic factors appear to be able to modify TDP-43-induced degeneration [50,51,52,53][50][51][52][53]. For example, using a mosaic genetic screen to study motor neuron degeneration in the Drosophila leg, Sreedharan et al. identified three factors, namely, sgg/GSK3, hat-trick, and xmas-2, needed to mediate TDP-43Q331K toxicity [50]. Interestingly, they noted that the manipulation of these three modifiers did not rescue Wallerian degeneration, another neurodegenerative but TDP-43-independent disease [50]. Among these three proteins, Shaggy (sgg), probably a downstream target of TDP-43, suppressed TDP-43 toxicity without reducing its expression [50]. In parallel, previous screening studies supported a mechanistic link between TDP-43 and Glycogen Synthase Kinase 3 (GSK3). They reported that TDP-43 activates GSK3, while GSK3 inhibition reduces TDP-43 aggregation [54,55][54][55]. Finally, the loss of the two other factors, xmas-2 or hat-trick, implicated in chromatin remodeling and RNA export, affected TDP-43 post-transcriptionally, resulting in a reduction in TDP-43 protein level [50]. In addition to this evidence, two recent studies using the suppressor screen techniques have led to the identifications of other epigenetic modifiers able to contrast TDP-43 neurotoxicity. In these two studies, human TDP-43 was over-expressed in a subset of Drosophila photoreceptor neurons [51], motor neurons, or glial cells [52]. Then, using a combination of shRNA or CRISPR/Cas9 knockdown (KD) screens, these cells were used to identify suppressors of TDP-43 neurotoxicity. Specifically, Azpurua and colleagues used an age-dependent neurobehavioral defect as a primary readout [52], while Berson and colleagues used the red eye degeneration readout [51]. Numerous genes implicated in nucleocytoplasmic transport or in pathways that are deregulated in TDP-43-related neurodegeneration were identified among the glial and motoneuronal TDP-43 suppressors of toxicity. However TDP-43-phenotype suppressors were principally composed of chromatin remodeling and basal transcription machinery factors. In particular, 25% of them were chromatin remodelers. Seven out of eight of these factors promote open chromatin as part of the Trithorax and SWI/SNF (Brahma) complexes, most of them with human known homologues: e(y)3/PHF10; polybromo/BAF180; ash1/ASH1L; enok/KAT6A; br/-; Br140/BRPF1; and mor/BAF170S/MARCC2. The remaining, the Chromodomain-helicase-DNA binding protein 1 (Chd1) with two human homologues, CHD1/CHD2, is an ATPase involved in the remodeling and assembly of chromatin [52]. Further analyses showed that TDP-43 can physically interact with fly Chd1 and human CHD2, impeding their recruitment onto chromatin. Interestingly, both proteins were clearly observed in the chromatin fractions but the Chd1-TDP-43 interaction did not take place on chromatin; rather, it was specifically observed in the cell-soluble fractions both in Drosophila and in human HEK293 cells [55]. By hijacking Chd1, overexpression of TDP-43 resulted in the impairment of correct nucleosome clearing from the gene body of a specific set of stress-protecting genes, preventing their activation [51]. The Chd1-TDP-43 interaction axis might therefore be one way by which the upregulation of TDP-43 sensitizes cells to various stress. Importantly, according to the different brain cells investigated, it was also observed that Chd1 KD could cause an opposite effect on TDP-43 overexpression-mediated toxicity; in particular, instead of counteracting the effects of TDP-43 overexpression upon motor neurons KD, an exacerbation of toxicity was observed upon glial and photoreceptor neurons KD [51,52][51][52]. These data highlight the importance of the cellular context for mediating TDP-43 activity, an important parameter that has been recently observed for TDP-43 pre-mRNA splicing regulatory properties as well [56]. Changing the model organism to C. elegans, it was found that the TDP-43 homolog, TDP-1, could regulate the chromatin localization of another chromatin remodeler, HPL-2, the heterochromatin protein 1 homolog [53]. Direct interaction was found to occur between these two proteins, both in the presence and absence of RNA. In this study, it was shown that TDP-1 facilitates HPL-2 association with active genes to maintain mRNA abundance. In addition, chromatin immunoprecipitation (ChIP) experiments indicated that TDP-1 is present at most of the HPL-2 peaks on chromatin in this organism. Specifically, loss of TDP-1 decreased most of the HPL-2 peaks where (AC)n and (AG)n binding motifs were present. These regions were located predominantly in intronic regions (71%) and promoters (20%) with levels of corresponding RNA decreasing in a tdp-1 mutant worm [53]. As a side note, the intronic localization of TDP-1 on DNA could be related to the propensity of TDP-1/TDP-43 orthologs to bind pre-mRNA chiefly within introns, as previously demonstrated in multiple organisms [34,35][34][35]. At the genome-wide level instead, TDP-43 was found to enrich particularly at promoter regulatory regions, will review in the next sections. Additional evidence of TDP-43 function in chromatin remodeling and its relevance to neurodegenerative diseases comes from the study of nBAFs proteins in cultured mouse motor neurons expressing ALS-linked mutant (G418C and A315T) human TDP-43 [57]. The Brahma-related gene 1 (Brg1)-associated factor (nBAF) chromatin-remodeling complex is critical for neuronal differentiation, dendritic extension, and synaptic function. In this study, the authors showed that nBAF subunits were lost in cultured mouse motor neurons expressing both mutants of human TDP-43. The decrease in nuclear Brg1, BAF53b, and CREST was observed when either mutant was expressed, but also when WT human TDP-43 protein expression was shifted to neuronal cytoplasmic inclusions, thus suggesting TDP-43 as a positive regulator of nBAF expression. In agreement with this conclusion, when co-expressed with mutant TDP-43, the presence of Brg1 delayed the induced dendritic attrition [57]. These data indicate that nuclear loss of TDP-43 can lead to a decrease in nBAF subunits production, either because of a transcriptional repression mechanism or following a defect in RNA processing, potentially leading to RNA nuclear retention, such as the one observed for Brg1 mRNA [57]. Nonetheless, it was interesting to observe that the depletion of nBAF subunits and the delayed attrition upon Brg1 co-expression were not unique to TDP-43; indeed, they were observed also for ALS-linked FUS mutants, and loss of nBAF subunits has also been reported to occur in spinal motor neurons of familial ALS (fALS) and sporadic ALS (sALS) patients with C9orf72 GC expansion (C9ALS) or sALS without mutations in common ALS-linked genes [57]. The contribution of TDP-43, and especially of its ALS-related mutants to more global epigenome alteration, was also recently tested in the human neuroblastoma SH-SY5Y cell line, together with other ALS-causative proteins, SOD1 and FUS [58]. In this work, the authors investigated four modifications on histone H3 tail associated with either transcriptional activation: (i) H3 serine 10 phosphorylation and lysine 14 acetylation (H3S10Ph-K14Ac); and (ii) H3 lysine 4 dimethylation (H3K4me2); or with transcriptional repression marks: (iii) H3 trimethylation of K9H3K9me3; and (iv) DNA methylation. Recombinant adenoviral expression of WT or ALS-related mutants of either TDP-43, SOD1 or FUS proteins all triggered a dose-dependent decrease in cell vitality. However, statistically significant differences in epigenetic marks were limited and specific to the TDP-43 genotype. In particular, a significant decrease in global H3S10Ph-K14Ac was observed for TDP-43M337V, whereas TDP-43WT overexpression led to a significant increase in H3K9me3. On the contrary, no relevant global losses or gains of these epigenetic marks were observed for the TDP-43A382T mutant [58]. In line with these findings, the fly Chd1-TDP-43 interaction study previously mentioned was part of a broad in vivo RNAi screen to search for TDP-43 toxicity modifiers [51]. This screen investigated a total of 84 genes related to various aspects of chromatin biology, including histone methyltransferases (HMTs), demethylases (HDMs), acetyltransferases (HATs), and deacetylases (HDACs), as well as associated factors, histones, and chromatin remodelers. [51]. In addition to Chd1, it allowed for the identification of an additional 4 ‘‘strong’’ and 27 ‘‘mild’’ modifiers, both enhancers and suppressors of TDP-43-mediated eye degeneration. Most of them converged on the conclusion that the TDP-43-mediated toxicity is associated to H3K4me3-linked aberrantly closed chromatin. The modulation of genes that alter other histone methylation marks (repressive H3K27, active-gene body H3K36, or H3K79) mostly had no effect on TDP-43 toxicity [51], with the exception of H3K9 HMT Su(var)3-9. The suppression in flies of Su(var)3-9—that is, the homolog of human SUV39H1—diminished TDP-43-induced toxicity [51]. These observations support the finding by Masala et al. of an aberrant increase in H3K9me3 modification upon ectopic TDP-43WT expression [58]. Note that this effect was not reported for the suppression of G9a, the other well-known H3K9 HMT [51]. Finally, two HDACs, HDAC1 and HDAC6, have been shown to influence and to be influenced by TDP-43, respectively. Thus, it has been shown that the silencing of both HDAC1 in SH-SY5Y and its fly ortholog Rpd3 in Drosophila is able to mitigate the toxic effect induced by TDP-43 expression [48]. Notably, this effect is possibly due to a direct modification of TDP-43 acetylation and consequent cellular localization and functional modulations, notably upon stress (see Section 2.2: TDP-43 and Local/Specific Gene Transcriptional Regulation). In 2010, two peer-reviewed studies showed that TDP-43 was able to bind HDAC6 mRNA, regulating both its mRNA and protein expression in neuronal and non-neuronal cell lines [59,60][59][60]. In one of these studies, Tibbetts’s group demonstrated that this interaction was also mediated by FUS/TLS, which was able to form protein complexes and to share overlapping HDAC6 binding sites with TDP-43 [60]. Conversely, HDAC6 was later shown to exert a deacetylation activity on TDP-43. Indeed, HDAC6 mediated the removal of TDP-43 acetylation at the residues Lys-145 and Lys-192, induced by the CPB acetyltransferase. This was found to decrease the cytoplasmic TDP-43 accumulation in otherwise normal cellular conditions [61]. On the contrary, the formation of TDP-43 aggregates that was induced in case of strong oxidative stress promoted by arsenite could not be deacetylated by HDAC6 despite its interaction with TDP-43, which, overall, contributed to the accumulation of mature aggregates of TDP-43 [61]. In 2020, the relationship between TDP-43 and HDAC6 was further analyzed by Lee and collaborators [62]. They found that the overexpression of HDAC6 in a Drosophila model of TDP-43 proteinopathy reduced the amount of insoluble poly-ubiquitinated proteins and ameliorated the lifespan and climbing defects associated with the overexpression of both TDP-43 and Ataxin-2 (ATXN2). These results indicated that HDAC6 could modulate, albeit in a non-enzymatic manner, the TDP-43 activity via the autophagy–lysosome pathway (ALP) [62]. At the level of gene expression, substantial alterations were observed in the cortices of transgenic mice expressing inducible WT or mutant hTDP-43 lacking the nuclear localization signal (tTA/TDPΔNLS). These alterations appeared even before the onset of significant gliosis and neuronal cell loss [63]. Despite both human TDP-43 transgenes downregulating the endogenous mTDP-43 (by the well-known phenomenon of TDP-43 autoregulation (see specific section)), the mutant lacking the nuclear localization signal showed the most profound changes in gene expression. Among the many processes that were altered in these mice, “DNA–protein complex assembly” pathway was particularly affected and harbored genes coding for major nucleosome proteins. Specifically, many histone variants (H2bp, H3d, H4a/H4b/H4c, and H4h) and several nucleosome assembly protein-1-like1 (NAP1L1) genes were found. While the histone variants were all upregulated, the NAP1L1 genes, on the contrary, were all downregulated [63]. Although these data were obtained using microarray, further RNA-seq analyses on the same model confirmed the alteration in transcription-related pathways and histone transcript levels [64]. In particular, it was observed that Med20, an essential component of the transcription-regulating Mediator complex, and Usp49, a histone H2B deubiquitinase which regulates splicing, were differentially spliced. In parallel, the canonical Histone Hist1h3 and Hist1h4 mRNAs were aberrantly polyadenylated, while at least 10 out of 15 variant histones were slightly but significantly downregulated in the TDPΔNLS bigenic mice [64]. In particular, enhanced cytoplasmic expression of TDP-43 downregulated histone 3′ UTR processing genes, notably Snrpe and Snrpd3, and a similar trend was observed for Lsm1l [64], thus further sustaining a role for TDP-43 in histone transcripts regulation. To relate these findings to the human pathological condition, it is now known that not all cells in the brain of a patient present a reduced load of nuclear TDP-43, and the transcriptome of these cellular populations was recently investigated [65]. To achieve this, Liu et al. successfully separated diseased neuronal nuclei without TDP-43 from nuclei retaining nuclear TDP-43 in a post-mortem FTD and FTD–ALS human brain by combining subcellular fractionation and fluorescent-activated cell sorting (FACS) [65]. Subsequent transcriptome analysis has revealed abundant changes in gene expression associated with loss of TDP-43. In keeping with results obtained from the various animal models, the data from this human material confirmed that many altered genes were involved in histone processing. Furthermore, DNA damage and repair genes were found enriched in addition to genes affecting proteostasis, RNA processing, and nucleocytoplasmic transport. In particular, it was noted that a cluster of 10 altered genes, namely, HUWE1, YY1, MORF4L2, HMGN1, PRKDC, UIMC1, POLB, SFPQ, MSH3, and XRCC5/Ku70, were part of a DNA repair module [65]. DNA methylation is another major epigenetic modification, acting on DNA itself, rather than on the chromatin or nucleosomal proteins wrapped around it. At the biological level, DNA methylation is established via DNA methyltransferases (DNMTs) and is passively erased during DNA replication or, as can be more relevant for neuronal cells, by active replication-independent mechanisms involving oxidations steps mediated by the ten-eleven translocation (TET) enzymes and base excision repair [66,67,68][66][67][68]. DNA methylation in mammals mostly takes place at cytosines (5mC) in the cytosine–guanine dinucleotide context (CpG), but 5mCpH (CpA, CpT, CpC) are also found in the adult mammalian brain [69]. The majority of the CpG are methylated in mammals, with dense CpG islands often unmethylated. CpG islands generally lie in the genes’ regulatory regions and impact transcription. CpG methylation generally has a repressive function, notably controlling promoter activation, but it can also regulate splicing and DNA stability [66,67,70,71][66][67][70][71]. On the other hand, its first oxidized state, the hydroxymethylated C (5hmC), positively influences gene expression, notably in the human brain [72]. No relevant changes in global DNA methylation were observed by Masala et al. in the human neuroblastoma SH-SY5Y cell line overexpressing WT or mutant ALS-linked proteins, including TDP-43, as cited above [58]; however, the brains of ALS patients show a different trend; in fact, altered DNA methylation has been recently observed to occur in human post-mortem CNS tissues from ALS patients using immunohistochemistry. It consisted of higher levels of 5mC and h5mC in the residual lower motor neurons of both sALS and C9ALS compared to the same region in controls [73]. A significantly lower number of neurons with detectable 5mC (mean about 28% vs. >73%) and 5hmC (mean about 51% vs. >87%) was found among neuronal subpopulations with pathological nuclear TDP-43 loss (10% of neurons) compared to those with normal nuclear TDP-43, therefore linking TPD-43 nuclear loss to loss of DNA methylation (despite the direction of causation remaining unknown). Overall, these findings could be connected to differential DNA methylation of several hundreds of genes in ALS spinal cord motor neurons, mostly involved in RNA processing and splicing [73]. Very recently, Catanese and colleagues used multi-omics and machine learning to question the transcriptional, epigenetic, and mutational aspects of heterogeneous human IPSCs-derived motor neurons holding mutants of either C9orf72, TARDBP, SOD1, or FUS, as well as datasets from patients’ biopsies [74]. Analysis of both transcriptome and methylation data resulted in different patterns characterizing the different ALS mutations. Thus, several thousands of DMRs were identified in the ALS sub-group as compared to control, yet a fraction (123 hypermethylated, 179 hypomethylated DMRs) was common to all subgroups, and partially overlapped with the TARDBP mutations (G298S and N390D)-holding subgroup [74]. These results also highlight a deep heterogeneity within the different ALS subtypes on the epigenetic level. Analysis of the DMR-related biological processes, however, indicated that epigenetic abnormalities among ALS iPSCs MNs all contribute to the synaptic alterations (downregulations) observed in all the related transcriptomes, although different sets of synaptic genes were hinted depending on the ALS-related mutation. Nonetheless, all the ALS iPCS-derived MNs displayed upregulation of acetylcholine receptor-binding genes in conjunction with a hypo-methylation of their promoters, notably LY6E, LY6H, and PSCA [74]. In addition, proteomic analysis of proteins co-purifying with TDP-43 in mice brain nuclear extracts has previously identified methyl CpG-binding protein 2 (MeCP2) as an interactor of TDP-43 [75]. MeCP2 is a protein whose defects are responsible for the degenerative Rett Syndrome pathology that binds mC and hmC not only in the CpG context. Interestingly, MeCP2 appears to be implicated in several regulatory contexts similar to TDP-43 (genes and TE transcription and RNA splicing, chromatin loop organization, and heterochromatin structure) [76]. Cell cycle alterations have also been reported following TDP-43 suppression. In two recent publications, TDP-43 activity was linked to sister chromatid cohesion through the splicing regulation of a cohesin complex subunit, namely, Stromal Antigen 2 (STAG2). In particular, depletion of TDP-43 in HeLa and neuroblastoma cell lines upregulated STAG2 exon 30b inclusion [77,78][77][78]. According to those data, cell accumulation was observed in G2/S phase, further supporting the role of TDP-43 in multiple processes involving genome remodeling. Finally, genes related to transcriptional machinery constitute another broad category of TDP-43-phenotype suppressors that have been identified thanks to several screening techniques. These genes include the transcription elongation factor, Su(Tpl) [52[52][79],79], which aberrantly expresses small nucleolar RNAs in TDP-43 pathology [79], TAF1, and e(y)1 orthologs of the mammalian TAF1 and TAF9 transcription factors, members of the TFIID initiation complex, and also Tombola involved in the transcriptional activation of the male germline during meiosis [79]. Specifically, no less than eight genes coding for subunits of the Mediator (Med) complex, mediating RNA-polymerase interaction with transcription factors, were identified by Azpurua et al. [52]. The alteration of another Med subunit, Med20, was identified in mice cortices upon TPD-43 manipulations [64], thus reinforcing a potential role for TDP-43 in gene transcriptional regulation on chromatin. Taken together, these studies on chromatin factors interacting with, modified by, or phenotypically rescuing TDP-43, indicate a potentially important role of TDP-43 as an epigenetic regulator with a high capacity for modulating chromatin, transcriptional processes, and DNA damage/repair pathways. A synthesis of the identified factors can be found in Table 1.
Table 1.
Chromatin and transcription factors directly or indirectly regulated by TDP-43 and modifying TDP-43-induced toxicity.

References

  1. Berson, A.; Nativio, R.; Berger, S.L.; Bonini, N.M. Epigenetic Regulation in Neurodegenerative Diseases. Trends Neurosci. 2018, 41, 587–598.
  2. Bennett, S.A.; Tanaz, R.; Cobos, S.N.; Torrente, M.P. Epigenetics in amyotrophic lateral sclerosis: A role for histone post-translational modifications in neurodegenerative disease. Transl. Res. 2019, 204, 19–30.
  3. Basavarajappa, B.S.; Subbanna, S. Histone Methylation Regulation in Neurodegenerative Disorders. Int. J. Mol. Sci. 2021, 22, 4654.
  4. Bodai, L. Epigenetic Regulation in Neurodegeneration Disease. Int. J. Mol. Sci. 2022, 23, 6185.
  5. Li, C.; Ren, J.; Zhang, M.; Wang, H.; Yi, F.; Wu, J.; Tang, Y. The heterogeneity of microglial activation and its epigenetic and non-coding RNA regulations in the immunopathogenesis of neurodegenerative diseases. Cell. Mol. Life Sci. 2022, 79, 511.
  6. Sharma, R.; Sharma, S.; Thakur, A.; Singh, A.; Singh, J.; Nepali, K.; Liou, J.P. The Role of Epigenetic Mechanisms in Autoimmune, Neurodegenerative, Cardiovascular, and Imprinting Disorders. Mini Rev. Med. Chem. 2022, 22, 1977–2011.
  7. Neumann, M.; Sampathu, D.M.; Kwong, L.K.; Truax, A.C.; Micsenyi, M.C.; Chou, T.T.; Bruce, J.; Schuck, T.; Grossman, M.; Clark, C.M.; et al. Ubiquitinated TDP-43 in Frontotemporal Lobar Degeneration and Amyotrophic Lateral Sclerosis. Science 2006, 314, 130–133.
  8. Arai, T.; Hasegawa, M.; Akiyama, H.; Ikeda, K.; Nonaka, T.; Mori, H.; Mann, D.; Tsuchiya, K.; Yoshida, M.; Hashizume, Y.; et al. TDP-43 is a component of ubiquitin-positive tau-negative inclusions in frontotemporal lobar degeneration and amyotrophic lateral sclerosis. Biochem. Biophys. Res. Commun. 2006, 351, 602–611.
  9. Prasad, A.; Bharathi, V.; Sivalingam, V.; Girdhar, A.; Patel, B.K. Molecular Mechanisms of TDP-43 Misfolding and Pathology in Amyotrophic Lateral Sclerosis. Front. Mol. Neurosci. 2019, 12, 25.
  10. Schwab, C.; Arai, T.; Hasegawa, M.; Yu, S.; McGeer, P.L. Colocalization of Transactivation-Responsive DNA-Binding Protein 43 and Huntingtin in Inclusions of Huntington Disease. J. Neuropathol. Exp. Neurol. 2008, 67, 1159–1165.
  11. Sanchez, I.I.; Nguyen, T.B.; England, W.E.; Lim, R.G.; Vu, A.Q.; Miramontes, R.; Byrne, L.M.; Markmiller, S.; Lau, A.L.; Orellana, I.; et al. Huntington’s disease mice and human brain tissue exhibit increased G3BP1 granules and TDP43 mislocalization. J. Clin. Investig. 2021, 131, 140723.
  12. Johnson, V.E.; Stewart, W.; Trojanowski, J.Q.; Smith, D.H. Acute and chronically increased immunoreactivity to phosphorylation-independent but not pathological TDP-43 after a single traumatic brain injury in humans. Acta Neuropathol. 2011, 122, 715–726.
  13. Gao, F.; Hu, M.; Zhang, J.; Hashem, J.; Chen, C. TDP-43 drives synaptic and cognitive deterioration following traumatic brain injury. Acta Neuropathol. 2022, 144, 187–210.
  14. De Boer, E.M.J.; Orie, V.K.; Williams, T.; Baker, M.R.; De Oliveira, H.M.; Polvikoski, T.; Silsby, M.; Menon, P.; van den Bos, M.; Halliday, G.M.; et al. TDP-43 proteinopathies: A new wave of neurodegenerative diseases. J. Neurol. Neurosurg. Psychiatry 2020, 92, 86–95.
  15. Ou, S.H.; Wu, F.; Harrich, D.; García-Martínez, L.F.; Gaynor, R.B. Cloning and characterization of a novel cellular protein, TDP-43, that binds to human immunodeficiency virus type 1 TAR DNA sequence motifs. J. Virol. 1995, 69, 3584–3596.
  16. Buratti, E.; Baralle, F.E. Characterization and Functional Implications of the RNA Binding Properties of Nuclear Factor TDP-43, a Novel Splicing Regulator ofCFTR Exon 9. J. Biol. Chem. 2001, 276, 36337–36343.
  17. Buratti, E.; Dörk, T.; Zuccato, E.; Pagani, F.; Romano, M.; Baralle, F.E. Nuclear factor TDP-43 and SR proteins promote in vitro and in vivo CFTR exon 9 skipping. EMBO J. 2001, 20, 1774–1784.
  18. Ayala, Y.M.; Pantano, S.; D’Ambrogio, A.; Buratti, E.; Brindisi, A.; Marchetti, C.; Romano, M.; Baralle, F.E. Human, Drosophila, and C. elegans TDP43: Nucleic Acid Binding Properties and Splicing Regulatory Function. J. Mol. Biol. 2005, 348, 575–588.
  19. Cassel, J.A.; Blass, B.E.; Reitz, A.B.; Pawlyk, A.C. Development of a Novel Nonradiometric Assay for Nucleic Acid Binding to TDP-43 Suitable for High-Throughput Screening Using AlphaScreen® Technology. J. Biomol. Screen. 2010, 15, 1099–1106.
  20. Kuo, P.-H.; Chiang, C.-H.; Wang, Y.-T.; Doudeva, L.G.; Yuan, H.S. The crystal structure of TDP-43 RRM1-DNA complex reveals the specific recognition for UG- and TG-rich nucleic acids. Nucleic Acids Res. 2014, 42, 4712–4722.
  21. Qin, H.; Lim, L.-Z.; Wei, Y.; Song, J. TDP-43 N terminus encodes a novel ubiquitin-like fold and its unfolded form in equilibrium that can be shifted by binding to ssDNA. Proc. Natl. Acad. Sci. USA 2014, 111, 18619–18624.
  22. Furukawa, Y.; Suzuki, Y.; Fukuoka, M.; Nagasawa, K.; Nakagome, K.; Shimizu, H.; Mukaiyama, A.; Akiyama, S. A molecular mechanism realizing sequence-specific recognition of nucleic acids by TDP-43. Sci. Rep. 2016, 6, 20576.
  23. Mitra, J.; Guerrero, E.N.; Hegde, P.M.; Liachko, N.F.; Wang, H.; Vasquez, V.; Gao, J.; Pandey, A.; Taylor, J.P.; Kraemer, B.C.; et al. Motor neuron disease-associated loss of nuclear TDP-43 is linked to DNA double-strand break repair defects. Proc. Natl. Acad. Sci. USA 2019, 116, 4696–4705.
  24. Ayala, Y.M.; Zago, P.; D’Ambrogio, A.; Xu, Y.-F.; Petrucelli, L.; Buratti, E.; Baralle, F.E. Structural determinants of the cellular localization and shuttling of TDP-43. J. Cell Sci. 2008, 121, 3778–3785.
  25. Alami, N.H.; Smith, R.B.; Carrasco, M.A.; Williams, L.A.; Winborn, C.S.; Han, S.S.; Kiskinis, E.; Winborn, B.; Freibaum, B.D.; Kanagaraj, A.; et al. Axonal Transport of TDP-43 mRNA Granules Is Impaired by ALS-Causing Mutations. Neuron 2014, 81, 536–543.
  26. Briese, M.; Saal-Bauernschubert, L.; Lüningschrör, P.; Moradi, M.; Dombert, B.; Surrey, V.; Appenzeller, S.; Deng, C.; Jablonka, S.; Sendtner, M. Loss of Tdp-43 disrupts the axonal transcriptome of motoneurons accompanied by impaired axonal translation and mitochondria function. Acta Neuropathol. Commun. 2020, 8, 116.
  27. Nagano, S.; Jinno, J.; Abdelhamid, R.F.; Jin, Y.; Shibata, M.; Watanabe, S.; Hirokawa, S.; Nishizawa, M.; Sakimura, K.; Onodera, O.; et al. TDP-43 transports ribosomal protein mRNA to regulate axonal local translation in neuronal axons. Acta Neuropathol. 2020, 140, 695–713.
  28. Nagano, S.; Araki, T. Axonal Transport and Local Translation of MRNA in Amyotrophic Lateral Sclerosis. In Amyotrophic Lateral Sclerosis; Araki, T., Ed.; Exon Publications: Brisbane, Australia, 2021; ISBN 978-0-645-00177-8.
  29. Winton, M.J.; Igaz, L.M.; Wong, M.M.; Kwong, L.K.; Trojanowski, J.Q.; Lee, V.M.-Y. Disturbance of Nuclear and Cytoplasmic TAR DNA-binding Protein (TDP-43) Induces Disease-like Redistribution, Sequestration, and Aggregate Formation. J. Biol. Chem. 2008, 283, 13302–13309.
  30. Johnson, B.S.; Snead, D.; Lee, J.J.; McCaffery, J.M.; Shorter, J.; Gitler, A.D. TDP-43 Is Intrinsically Aggregation-prone, and Amyotrophic Lateral Sclerosis-linked Mutations Accelerate Aggregation and Increase Toxicity. J. Biol. Chem. 2009, 284, 20329–20339.
  31. Berning, B.A.; Walker, A.K. The Pathobiology of TDP-43 C-Terminal Fragments in ALS and FTLD. Front. Neurosci. 2019, 13, 335.
  32. Buratti, E.; Baralle, F.E. The multiple roles of TDP-43 in pre-mRNA processing and gene expression regulation. RNA Biol. 2010, 7, 420–429.
  33. Buratti, E.; De Conti, L.; Stuani, C.; Romano, M.; Baralle, M.; Baralle, F. Nuclear factor TDP-43 can affect selected microRNA levels. FEBS J. 2010, 277, 2268–2281.
  34. Polymenidou, M.; Lagier-Tourenne, C.; Hutt, K.R.; Huelga, S.C.; Moran, J.; Liang, T.Y.; Ling, S.-C.; Sun, E.; Wancewicz, E.; Mazur, C.; et al. Long pre-mRNA depletion and RNA missplicing contribute to neuronal vulnerability from loss of TDP-43. Nat. Neurosci. 2011, 14, 459–468.
  35. Tollervey, J.R.; Curk, T.; Rogelj, B.; Briese, M.; Cereda, M.; Kayikci, M.; König, J.; Hortobágyi, T.; Nishimura, A.L.; Župunski, V.; et al. Characterizing the RNA targets and position-dependent splicing regulation by TDP-43. Nat. Neurosci. 2011, 14, 452–458.
  36. Kawahara, Y.; Mieda-Sato, A. TDP-43 promotes microRNA biogenesis as a component of the Drosha and Dicer complexes. Proc. Natl. Acad. Sci. USA 2012, 109, 3347–3352.
  37. Rathore, A.; Iketani, S.; Wang, P.; Jia, M.; Sahi, V.; Ho, D.D. CRISPR-based gene knockout screens reveal deubiquitinases involved in HIV-1 latency in two Jurkat cell models. Sci. Rep. 2020, 10, 5350.
  38. Acharya, K.K.; Govind, C.K.; Shore, A.N.; Stoler, M.H.; Reddi, P.P. cis-Requirement for the maintenance of round spermatid-specific transcription. Dev. Biol. 2006, 295, 781–790.
  39. Abhyankar, M.M.; Urekar, C.; Reddi, P.P. A Novel CpG-free Vertebrate Insulator Silences the Testis-specific SP-10 Gene in Somatic Tissues: Role for TDP-43 in Insulator Function. J. Biol. Chem. 2007, 282, 36143–36154.
  40. Lalmansingh, A.S.; Urekar, C.J.; Reddi, P.P. TDP-43 Is a Transcriptional Repressor: The Testis-Specific Mouse Acrv1 Gene Is a TDP-43 Target in Vivo. J. Biol. Chem. 2011, 286, 10970–10982.
  41. Schwenk, B.M.; Hartmann, H.; Serdaroglu, A.; Schludi, M.H.; Hornburg, D.; Meissner, F.; Orozco, D.; Colombo, A.; Tahirovic, S.; Michaelsen, M.; et al. TDP-43 loss of function inhibits endosomal trafficking and alters trophic signaling in neurons. EMBO J. 2016, 35, 2350–2370.
  42. Militello, G.; Hosen, M.R.; Ponomareva, Y.; Gellert, P.; Weirick, T.; John, D.; Hindi, S.M.; Mamchaoui, K.; Mouly, V.; Döring, C.; et al. A novel long non-coding RNA Myolinc regulates myogenesis through TDP-43 and Filip1. J. Mol. Cell Biol. 2018, 10, 102–117.
  43. Keihani, S.; Kluever, V.; Mandad, S.; Bansal, V.; Rahman, R.; Fritsch, E.; Gomes, L.C.; Gärtner, A.; Kügler, S.; Urlaub, H.; et al. The long noncoding RNA neuroLNC regulates presynaptic activity by interacting with the neurodegeneration-associated protein TDP-43. Sci. Adv. 2019, 5, eaay2670.
  44. Pandya-Jones, A.; Markaki, Y.; Serizay, J.; Chitiashvili, T.; Leon, W.R.M.; Damianov, A.; Chronis, C.; Papp, B.; Chen, C.-K.; McKee, R.; et al. A protein assembly mediates Xist localization and gene silencing. Nature 2020, 587, 145–151.
  45. Li, P.; Ruan, X.; Yang, L.; Kiesewetter, K.; Zhao, Y.; Luo, H.; Chen, Y.; Gucek, M.; Zhu, J.; Cao, H. A Liver-Enriched Long Non-Coding RNA, lncLSTR, Regulates Systemic Lipid Metabolism in Mice. Cell Metab. 2015, 21, 455–467.
  46. Murata, H.; Hattori, T.; Maeda, H.; Takashiba, S.; Takigawa, M.; Kido, J.; Nagata, T. Identification of transactivation-responsive DNA-binding protein 43 (TARDBP43; TDP-43) as a novel factor for TNF-α expression upon lipopolysaccharide stimulation in human monocytes. J. Periodontal Res. 2015, 50, 452–460.
  47. Suzuki, H.; Matsuoka, M. TDP-43 toxicity is mediated by the unfolded protein response-unrelated induction of C/EBP homologous protein expression. J. Neurosci. Res. 2012, 90, 641–647.
  48. Sanna, S.; Esposito, S.; Masala, A.; Sini, P.; Nieddu, G.; Galioto, M.; Fais, M.; Iaccarino, C.; Cestra, G.; Crosio, C. HDAC1 inhibition ameliorates TDP-43-induced cell death in vitro and in vivo. Cell Death Dis. 2020, 11, 369.
  49. Li, J.; Zhao, W.; Li, Q.; Huang, Z.; Shi, G.; Li, C. Long Non-Coding RNA H19 Promotes Porcine Satellite Cell Differentiation by Interacting with TDP43. Genes 2020, 11, 259.
  50. Sreedharan, J.; Neukomm, L.J.; Brown, R.H.; Freeman, M.R. Age-Dependent TDP-43-Mediated Motor Neuron Degeneration Requires GSK3, hat-trick, and xmas-2. Curr. Biol. 2015, 25, 2130–2136.
  51. Berson, A.; Sartoris, A.; Nativio, R.; Van Deerlin, V.; Toledo, J.B.; Porta, S.; Liu, S.; Chung, C.-Y.; Garcia, B.A.; Lee, V.M.-Y.; et al. TDP-43 Promotes Neurodegeneration by Impairing Chromatin Remodeling. Curr. Biol. 2017, 27, 3579–3590.e6.
  52. Azpurua, J.; El-Karim, E.G.; Tranquille, M.; Dubnau, J. A behavioral screen for mediators of age-dependent TDP-43 neurodegeneration identifies SF2/SRSF1 among a group of potent suppressors in both neurons and glia. PLoS Genet. 2021, 17, e1009882.
  53. Saldi, T.K.; Gonzales, P.; Garrido-Lecca, A.; Dostal, V.; Roberts, C.M.; Petrucelli, L.; Link, C.D. The Caenorhabditis elegans Ortholog of TDP-43 Regulates the Chromatin Localization of the Heterochromatin Protein 1 Homolog HPL-2. Mol. Cell. Biol. 2018, 38, e00668-17.
  54. Moujalled, D.; James, J.L.; Parker, S.J.; Lidgerwood, G.E.; Duncan, C.; Meyerowitz, J.; Nonaka, T.; Hasegawa, M.; Kanninen, K.M.; Grubman, A.; et al. Kinase Inhibitor Screening Identifies Cyclin-Dependent Kinases and Glycogen Synthase Kinase 3 as Potential Modulators of TDP-43 Cytosolic Accumulation during Cell Stress. PLoS ONE 2013, 8, e67433.
  55. Ambegaokar, S.S.; Jackson, G.R. Functional genomic screen and network analysis reveal novel modifiers of tauopathy dissociated from tau phosphorylation. Hum. Mol. Genet. 2011, 20, 4947–4977.
  56. Šušnjar, U.; Škrabar, N.; Brown, A.-L.; Abbassi, Y.; Phatnani, H.; NYGC ALS Consortium; Cortese, A.; Cereda, C.; Bugiardini, E.; Cardani, R.; et al. Cell environment shapes TDP-43 function with implications in neuronal and muscle disease. Commun. Biol. 2022, 5, 314.
  57. Tibshirani, M.; Zhao, B.; Gentil, B.J.; Minotti, S.; Marques, C.; Keith, J.; Rogaeva, E.; Zinman, L.; Rouaux, C.; Robertson, J.; et al. Dysregulation of chromatin remodelling complexes in amyotrophic lateral sclerosis. Hum. Mol. Genet. 2017, 26, 4142–4152.
  58. Masala, A.; Sanna, S.; Esposito, S.; Rassu, M.; Galioto, M.; Zinellu, A.; Carru, C.; Carrì, M.T.; Iaccarino, C.; Crosio, C. Epigenetic Changes Associated with the Expression of Amyotrophic Lateral Sclerosis (ALS) Causing Genes. Neuroscience 2018, 390, 1–11.
  59. Fiesel, F.C.; Voigt, A.; Weber, S.S.; Van den Haute, C.; Waldenmaier, A.; Görner, K.; Walter, M.; Anderson, M.L.; Kern, J.V.; Rasse, T.M.; et al. Knockdown of transactive response DNA-binding protein (TDP-43) downregulates histone deacetylase 6. EMBO J. 2010, 29, 209–221.
  60. Kim, S.H.; Shanware, N.P.; Bowler, M.J.; Tibbetts, R.S. Amyotrophic Lateral Sclerosis-associated Proteins TDP-43 and FUS/TLS Function in a Common Biochemical Complex to Co-regulate HDAC6 mRNA*. J. Biol. Chem. 2010, 285, 34097–34105.
  61. Cohen, T.J.; Hwang, A.W.; Restrepo, C.R.; Yuan, C.-X.; Trojanowski, J.Q.; Lee, V.M.Y. An acetylation switch controls TDP-43 function and aggregation propensity. Nat. Commun. 2015, 6, 5845.
  62. Lee, S.; Kwon, Y.; Kim, S.; Jo, M.; Jeon, Y.-M.; Cheon, M.; Lee, S.; Kim, S.R.; Kim, K.; Kim, H.-J. The Role of HDAC6 in TDP-43-Induced Neurotoxicity and UPS Impairment. Front. Cell Dev. Biol. 2020, 8, 581942.
  63. Igaz, L.M.; Kwong, L.K.; Lee, E.B.; Chen-Plotkin, A.; Swanson, E.; Unger, T.; Malunda, J.; Xu, Y.; Winton, M.J.; Trojanowski, J.Q.; et al. Dysregulation of the ALS-associated gene TDP-43 leads to neuronal death and degeneration in mice. J. Clin. Investig. 2011, 121, 726–738.
  64. Amlie-Wolf, A.; Ryvkin, P.; Tong, R.; Dragomir, I.; Suh, E.; Xu, Y.; Van Deerlin, V.M.; Gregory, B.D.; Kwong, L.K.; Trojanowski, J.Q.; et al. Transcriptomic Changes Due to Cytoplasmic TDP-43 Expression Reveal Dysregulation of Histone Transcripts and Nuclear Chromatin. PLoS ONE 2015, 10, e0141836.
  65. Liu, E.Y.; Russ, J.; Cali, C.P.; Phan, J.M.; Amlie-Wolf, A.; Lee, E.B. Loss of Nuclear TDP-43 Is Associated with Decondensation of LINE Retrotransposons. Cell Rep. 2019, 27, 1409–1421.e6.
  66. Bird, A. DNA methylation patterns and epigenetic memory. Genes Dev. 2002, 16, 6–21.
  67. Moore, L.D.; Le, T.; Fan, G. DNA Methylation and Its Basic Function. Neuropsychopharmacology 2013, 38, 23–38.
  68. Gavin, D.P.; Chase, K.A.; Sharma, R.P. Active DNA demethylation in post-mitotic neurons: A reason for optimism. Neuropharmacology 2013, 75, 233–245.
  69. Guo, J.U.; Su, Y.; Shin, J.H.; Shin, J.; Li, H.; Xie, B.; Zhong, C.; Hu, S.; Le, T.; Fan, G.; et al. Distribution, recognition and regulation of non-CpG methylation in the adult mammalian brain. Nat. Neurosci. 2013, 17, 215–222.
  70. Robertson, K.D.; Wolffe, A.P. DNA methylation in health and disease. Nat. Rev. Genet. 2000, 1, 11–19.
  71. Wu, H.; Coskun, V.; Tao, J.; Xie, W.; Ge, W.; Yoshikawa, K.; Li, E.; Zhang, Y.; Sun, Y.E. Dnmt3a-Dependent Nonpromoter DNA Methylation Facilitates Transcription of Neurogenic Genes. Science 2010, 329, 444–448.
  72. Wang, J.; Zhang, K.X.; Lu, G.Z.; Zhao, X.H. Research progress on 5hmC and TET dioxygenases in neurodevelopment and neurological diseases. Yi Chuan = Hered. 2017, 39, 1138–1149.
  73. Appleby-Mallinder, C.; Schaber, E.; Kirby, J.; Shaw, P.J.; Cooper-Knock, J.; Heath, P.R.; Highley, J.R. TDP43 proteinopathy is associated with aberrant DNA methylation in human amyotrophic lateral sclerosis. Neuropathol. Appl. Neurobiol. 2020, 47, 61–72.
  74. Catanese, A.; Rajkumar, S.; Sommer, D.; Masrori, P.; Hersmus, N.; Van Damme, P.; Witzel, S.; Ludolph, A.; Ho, R.; Boeckers, T.M.; et al. Multiomics and Machine-Learning Identify Novel Transcriptional and Mutational Signatures in Amyotrophic Lateral Sclerosis. Brain J. Neurol. 2023, awad075.
  75. Sephton, C.F.; Cenik, C.; Kucukural, A.; Dammer, E.B.; Cenik, B.; Han, Y.; Dewey, C.M.; Roth, F.P.; Herz, J.; Peng, J.; et al. Identification of Neuronal RNA Targets of TDP-43-containing Ribonucleoprotein Complexes. J. Biol. Chem. 2011, 286, 1204–1215.
  76. Sharifi, O.; Yasui, D.H. The Molecular Functions of MeCP2 in Rett Syndrome Pathology. Front. Genet. 2021, 12, 624290.
  77. De Conti, L.; Akinyi, M.V.; Mendoza-Maldonado, R.; Romano, M.; Baralle, M.; Buratti, E. TDP-43 affects splicing profiles and isoform production of genes involved in the apoptotic and mitotic cellular pathways. Nucleic Acids Res. 2015, 43, 8990–9005.
  78. Appocher, C.; Mohagheghi, F.; Cappelli, S.; Stuani, C.; Romano, M.; Feiguin, F.; Buratti, E. Major hnRNP proteins act as general TDP-43 functional modifiers both in Drosophila and human neuronal cells. Nucleic Acids Res. 2017, 45, 8026–8045.
  79. Chung, C.-Y.; Berson, A.; Kennerdell, J.R.; Sartoris, A.; Unger, T.; Porta, S.; Kim, H.-J.; Smith, E.R.; Shilatifard, A.; Van Deerlin, V.; et al. Aberrant activation of non-coding RNA targets of transcriptional elongation complexes contributes to TDP-43 toxicity. Nat. Commun. 2018, 9, 4406.
  80. Swarup, V.; Phaneuf, D.; Dupré, N.; Petri, S.; Strong, M.; Kriz, J.; Julien, J.-P. Deregulation of TDP-43 in amyotrophic lateral sclerosis triggers nuclear factor κB–mediated pathogenic pathways. J. Exp. Med. 2011, 208, 2429–2447.
  81. Janssen, C.; Schmalbach, S.; Boeselt, S.; Sarlette, A.; Dengler, R.; Petri, S. Differential Histone Deacetylase mRNA Expression Patterns in Amyotrophic Lateral Sclerosis. J. Neuropathol. Exp. Neurol. 2010, 69, 573–581.
  82. Casafont, I.; Bengoechea, R.; Tapia, O.; Berciano, M.T.; Lafarga, M. TDP-43 localizes in mRNA transcription and processing sites in mammalian neurons. J. Struct. Biol. 2009, 167, 235–241.
  83. Swain, A.; Misulovin, Z.; Pherson, M.; Gause, M.; Mihindukulasuriya, K.; Rickels, R.A.; Shilatifard, A.; Dorsett, D. Drosophila TDP-43 RNA-Binding Protein Facilitates Association of Sister Chromatid Cohesion Proteins with Genes, Enhancers and Polycomb Response Elements. PLOS Genet. 2016, 12, e1006331.
  84. Davidovich, C.; Cech, T.R. The recruitment of chromatin modifiers by long noncoding RNAs: Lessons from PRC2. RNA 2015, 21, 2007–2022.
  85. Schuettengruber, B.; Bourbon, H.-M.; Di Croce, L.; Cavalli, G. Genome Regulation by Polycomb and Trithorax: 70 Years and Counting. Cell 2017, 171, 34–57.
  86. Laugesen, A.; Højfeldt, J.W.; Helin, K. Molecular Mechanisms Directing PRC2 Recruitment and H3K27 Methylation. Mol. Cell 2019, 74, 8–18.
  87. Owen, B.M.; Davidovich, C. DNA binding by polycomb-group proteins: Searching for the link to CpG islands. Nucleic Acids Res. 2022, 50, 4813–4839.
  88. Wang, X.; Goodrich, K.J.; Conlon, E.G.; Gao, J.; Erbse, A.H.; Manley, J.L.; Cech, T.R. C9orf72 and triplet repeat disorder RNAs: G-quadruplex formation, binding to PRC2 and implications for disease mechanisms. RNA 2019, 25, 935–947.
  89. Morera, A.A.; Ahmed, N.S.; Schwartz, J.C. TDP-43 regulates transcription at protein-coding genes and Alu retrotransposons. Biochim. Biophys. Acta Gene Regul. Mech. 2019, 1862, 194434.
  90. Maor-Nof, M.; Shipony, Z.; Lopez-Gonzalez, R.; Nakayama, L.; Zhang, Y.-J.; Couthouis, J.; Blum, J.A.; Castruita, P.A.; Linares, G.R.; Ruan, K.; et al. p53 is a central regulator driving neurodegeneration caused by C9orf72 poly(PR). Cell 2021, 184, 689–708.e20.
  91. Nie, X.; Xu, Q.; Xu, C.; Chen, F.; Wang, Q.; Qin, D.; Wang, R.; Gao, Z.; Lu, X.; Yang, X.; et al. Maternal TDP-43 interacts with RNA Pol II and regulates zygotic genome activation. Nat. Commun. 2023, 14, 4275.
  92. Hill, S.J.; Mordes, D.A.; Cameron, L.A.; Neuberg, D.S.; Landini, S.; Eggan, K.; Livingston, D.M. Two familial ALS proteins function in prevention/repair of transcription-associated DNA damage. Proc. Natl. Acad. Sci. USA 2016, 113, E7701–E7709.
  93. Guerrero, E.N.; Mitra, J.; Wang, H.; Rangaswamy, S.; Hegde, P.M.; Basu, P.; Rao, K.S.; Hegde, M.L. Amyotrophic lateral sclerosis-associated TDP-43 mutation Q331K prevents nuclear translocation of XRCC4-DNA ligase 4 complex and is linked to genome damage-mediated neuronal apoptosis. Hum. Mol. Genet. 2019, 28, 2459–2476.
  94. Kawaguchi, T.; Rollins, M.G.; Moinpour, M.; Morera, A.A.; Ebmeier, C.C.; Old, W.M.; Schwartz, J.C. Changes to the TDP-43 and FUS Interactomes Induced by DNA Damage. J. Proteome Res. 2019, 19, 360–370.
  95. Konopka, A.; Whelan, D.R.; Jamali, S.; Perri, E.; Shahheydari, H.; Toth, R.P.; Parakh, S.; Robinson, T.; Cheong, A.; Mehta, P.; et al. Impaired NHEJ repair in amyotrophic lateral sclerosis is associated with TDP-43 mutations. Mol. Neurodegener. 2020, 15, 51.
  96. Eldridge, M.J.G.; Hamon, M.A. Histone H3 deacetylation promotes host cell viability for efficient infection by Listeria monocytogenes. PLoS Pathog. 2021, 17, e1010173.
  97. Giannini, M.; Bayona-Feliu, A.; Sproviero, D.; Barroso, S.I.; Cereda, C.; Aguilera, A. TDP-43 mutations link Amyotrophic Lateral Sclerosis with R-loop homeostasis and R loop-mediated DNA damage. PLoS Genet. 2020, 16, e1009260.
  98. Sharma, S. Age-related nonhomologous end joining activity in rat neurons. Brain Res. Bull. 2007, 73, 48–54.
  99. Yu, A.M.; McVey, M. Synthesis-dependent microhomology-mediated end joining accounts for multiple types of repair junctions. Nucleic Acids Res. 2010, 38, 5706–5717.
  100. Eskandarian, H.A.; Impens, F.; Nahori, M.-A.; Soubigou, G.; Coppée, J.-Y.; Cossart, P.; Hamon, M.A. A Role for SIRT2-Dependent Histone H3K18 Deacetylation in Bacterial Infection. Science 2013, 341, 1238858.
  101. Chen, X.; Lu, W.; Wu, D. Sirtuin 2 (SIRT2): Confusing Roles in the Pathophysiology of Neurological Disorders. Front. Neurosci. 2021, 15, 614107.
  102. Yu, Z.; Fan, D.; Gui, B.; Shi, L.; Xuan, C.; Shan, L.; Wang, Q.; Shang, Y.; Wang, Y. Neurodegeneration-associated TDP-43 Interacts with Fragile X Mental Retardation Protein (FMRP)/Staufen (STAU1) and Regulates SIRT1 Expression in Neuronal Cells. J. Biol. Chem. 2012, 287, 22560–22572.
  103. Zhang, W.; Wu, H.; Yang, M.; Ye, S.; Li, L.; Zhang, H.; Hu, J.; Wang, X.; Xu, J.; Liang, A. SIRT1 inhibition impairs non-homologous end joining DNA damage repair by increasing Ku70 acetylation in chronic myeloid leukemia cells. Oncotarget 2015, 7, 13538–13550.
  104. Herskovits, A.Z.; Guarente, L. SIRT1 in Neurodevelopment and Brain Senescence. Neuron 2014, 81, 471–483.
  105. Singh, M.; Zhang, S.; Perez, A.M.; Lee, E.Y.C.; Lee, M.Y.W.T.; Zhang, D. POLDIP3: At the Crossroad of RNA and DNA Metabolism. Genes 2022, 13, 1921.
  106. Shiga, A.; Ishihara, T.; Miyashita, A.; Kuwabara, M.; Kato, T.; Watanabe, N.; Yamahira, A.; Kondo, C.; Yokoseki, A.; Takahashi, M.; et al. Alteration of POLDIP3 Splicing Associated with Loss of Function of TDP-43 in Tissues Affected with ALS. PLoS ONE 2012, 7, e43120.
  107. Wang, W.-Y.; Pan, L.; Su, S.C.; Quinn, E.J.; Sasaki, M.; Jimenez, J.C.; Mackenzie, I.R.A.; Huang, E.J.; Tsai, L.-H. Interaction of FUS and HDAC1 regulates DNA damage response and repair in neurons. Nat. Neurosci. 2013, 16, 1383–1391.
  108. Wang, H.; Guo, W.; Mitra, J.; Hegde, P.M.; Vandoorne, T.; Eckelmann, B.J.; Mitra, S.; Tomkinson, A.E.; Van Den Bosch, L.; Hegde, M.L. Mutant FUS causes DNA ligation defects to inhibit oxidative damage repair in Amyotrophic Lateral Sclerosis. Nat. Commun. 2018, 9, 3683.
  109. Stirling, P.C.; Hieter, P. Canonical DNA Repair Pathways Influence R-Loop-Driven Genome Instability. J. Mol. Biol. 2017, 429, 3132–3138.
  110. Cuartas, J.; Gangwani, L. R-loop Mediated DNA Damage and Impaired DNA Repair in Spinal Muscular Atrophy. Front. Cell. Neurosci. 2022, 16, 826608.
  111. Wood, M.; Quinet, A.; Lin, Y.-L.; Davis, A.A.; Pasero, P.; Ayala, Y.M.; Vindigni, A. TDP-43 dysfunction results in R-loop accumulation and DNA replication defects. J. Cell Sci. 2020, 133, jcs244129.
  112. Gong, D.; Wang, L.; Zhou, H.; Gao, J.; Zhang, W.; Zheng, P. Long noncoding RNA Lnc530 localizes on R-loops and regulates R-loop formation and genomic stability in mouse embryonic stem cells. Stem Cell Rep. 2023, 18, 952–968.
  113. Lander, E.S.; Linton, L.M.; Birren, B.; Nusbaum, C.; Zody, M.C.; Baldwin, J.; Devon, K.; Dewar, K.; Doyle, M.; FitzHugh, W.; et al. Initial sequencing and analysis of the human genome. Nature 2001, 409, 860–921.
  114. Bannert, N.; Kurth, R. The Evolutionary Dynamics of Human Endogenous Retroviral Families. Annu. Rev. Genom. Hum. Genet. 2006, 7, 149–173.
  115. Villesen, P.; Aagaard, L.; Wiuf, C.; Pedersen, F.S. Identification of endogenous retroviral reading frames in the human genome. Retrovirology 2004, 1, 32.
  116. Ueda, M.T.; Kryukov, K.; Mitsuhashi, S.; Mitsuhashi, H.; Imanishi, T.; Nakagawa, S. Comprehensive genomic analysis reveals dynamic evolution of endogenous retroviruses that code for retroviral-like protein domains. Mob. DNA 2020, 11, 29.
  117. Moyes, D.; Griffiths, D.J.; Venables, P.J. Insertional polymorphisms: A new lease of life for endogenous retroviruses in human disease. Trends Genet. 2007, 23, 326–333.
  118. Wildschutte, J.H.; Williams, Z.H.; Montesion, M.; Subramanian, R.P.; Kidd, J.M.; Coffin, J.M. Discovery of unfixed endogenous retrovirus insertions in diverse human populations. Proc. Natl. Acad. Sci. USA 2016, 113, E2326–E2334.
  119. Li, W.; Lin, L.; Malhotra, R.; Yang, L.; Acharya, R.; Poss, M. A computational framework to assess genome-wide distribution of polymorphic human endogenous retrovirus-K In human populations. PLoS Comput. Biol. 2019, 15, e1006564.
  120. Dupressoir, A.; Lavialle, C.; Heidmann, T. From ancestral infectious retroviruses to bona fide cellular genes: Role of the captured syncytins in placentation. Placenta 2012, 33, 663–671.
  121. Kaneko-Ishino, T.; Ishino, F. The role of genes domesticated from LTR retrotransposons and retroviruses in mammals. Front. Microbiol. 2012, 3, 262.
  122. Naville, M.; Warren, I.; Haftek-Terreau, Z.; Chalopin, D.; Brunet, F.; Levin, P.; Galiana, D.; Volff, J.-N. Not so bad after all: Retroviruses and long terminal repeat retrotransposons as a source of new genes in vertebrates. Clin. Microbiol. Infect. 2016, 22, 312–323.
  123. Grandi, N.; Tramontano, E. Human Endogenous Retroviruses Are Ancient Acquired Elements Still Shaping Innate Immune Responses. Front. Immunol. 2018, 9, 2039.
  124. Mager, D.L.; Stoye, J.P. Mammalian Endogenous Retroviruses. Microbiol. Spectr. 2015, 3, 1079–1100.
  125. Brouha, B.; Schustak, J.; Badge, R.M.; Lutz-Prigge, S.; Farley, A.H.; Moran, J.V.; Kazazian, H.H. Hot L1s account for the bulk of retrotransposition in the human population. Proc. Natl. Acad. Sci. USA 2003, 100, 5280–5285.
  126. Dewannieux, M.; Heidmann, T. LINEs, SINEs and processed pseudogenes: Parasitic strategies for genome modeling. Cytogenet. Genome Res. 2005, 110, 35–48.
  127. Pavlíček, A.; Pačes, J.; Elleder, D.; Hejnar, J. Processed Pseudogenes of Human Endogenous Retroviruses Generated by LINEs: Their Integration, Stability, and Distribution. Genome Res. 2002, 12, 391–399.
  128. Mouse Genome Sequencing Consortium; Waterston, R.H.; Lindblad-Toh, K.; Birney, E.; Rogers, J.; Abril, J.F.; Agarwal, P.; Agarwala, R.; Ainscough, R.; Alexandersson, M.; et al. Initial sequencing and comparative analysis of the mouse genome. Nature 2002, 420, 520–562.
  129. Coufal, N.G.; Garcia-Perez, J.L.; Peng, G.E.; Yeo, G.W.; Mu, Y.; Lovci, M.T.; Morell, M.; O’Shea, K.S.; Moran, J.V.; Gage, F.H. L1 retrotransposition in human neural progenitor cells. Nature 2009, 460, 1127–1131.
  130. Della Valle, F.; Thimma, M.P.; Caiazzo, M.; Pulcrano, S.; Celii, M.; Adroub, S.A.; Liu, P.; Alanis-Lobato, G.; Broccoli, V.; Orlando, V. Transdifferentiation of Mouse Embryonic Fibroblasts into Dopaminergic Neurons Reactivates LINE-1 Repetitive Elements. Stem Cell Rep. 2020, 14, 60–74.
  131. Richardson, S.R.; Morell, S.; Faulkner, G.J. L1 Retrotransposons and Somatic Mosaicism in the Brain. Annu. Rev. Genet. 2014, 48, 1–27.
  132. Jern, P.; Sperber, G.O.; Blomberg, J. Use of Endogenous Retroviral Sequences (ERVs) and structural markers for retroviral phylogenetic inference and taxonomy. Retrovirology 2005, 2, 50.
  133. Groh, S.; Schotta, G. Silencing of endogenous retroviruses by heterochromatin. Cell. Mol. Life Sci. 2017, 74, 2055–2065.
  134. Geis, F.K.; Goff, S.P. Silencing and Transcriptional Regulation of Endogenous Retroviruses: An Overview. Viruses 2020, 12, 884.
  135. Goodier, J.L. Restricting retrotransposons: A review. Mob. DNA 2016, 7, 16.
  136. Savage, A.L.; Lopez, A.I.; Iacoangeli, A.; Bubb, V.J.; Smith, B.; Troakes, C.; Alahmady, N.; Koks, S.; Schumann, G.G.; Al-Chalabi, A.; et al. Frequency and methylation status of selected retrotransposition competent L1 loci in amyotrophic lateral sclerosis. Mol. Brain 2020, 13, 154.
  137. Li, W.; Jin, Y.; Prazak, L.; Hammell, M.; Dubnau, J. Transposable Elements in TDP-43-Mediated Neurodegenerative Disorders. PLoS ONE 2012, 7, e44099.
  138. Tam, O.H.; Rozhkov, N.V.; Shaw, R.; Kim, D.; Hubbard, I.; Fennessey, S.; Propp, N.; NYGC ALS Consortium; Fagegaltier, D.; Harris, B.T.; et al. Postmortem Cortex Samples Identify Distinct Molecular Subtypes of ALS: Retrotransposon Activation, Oxidative Stress, and Activated Glia. Cell Rep. 2019, 29, 1164–1177.e5.
  139. Kelley, D.R.; Hendrickson, D.G.; Tenen, D.; Rinn, J.L. Transposable elements modulate human RNA abundance and splicing via specific RNA-protein interactions. Genome Biol. 2014, 15, 537.
  140. Douville, R.; Liu, J.; Rothstein, J.; Nath, A. Identification of active loci of a human endogenous retrovirus in neurons of patients with amyotrophic lateral sclerosis. Ann. Neurol. 2011, 69, 141–151.
  141. Li, W.; Lee, M.-H.; Henderson, L.; Tyagi, R.; Bachani, M.; Steiner, J.; Campanac, E.; Hoffman, D.A.; von Geldern, G.; Johnson, K.; et al. Human endogenous retrovirus-K contributes to motor neuron disease. Sci. Transl. Med. 2015, 7, 307ra153.
  142. Ibba, G.; Piu, C.; Uleri, E.; Serra, C.; Dolei, A. Disruption by SaCas9 Endonuclease of HERV-Kenv, a Retroviral Gene with Oncogenic and Neuropathogenic Potential, Inhibits Molecules Involved in Cancer and Amyotrophic Lateral Sclerosis. Viruses 2018, 10, 412.
  143. Fuentes, D.R.; Swigut, T.; Wysocka, J. Systematic perturbation of retroviral LTRs reveals widespread long-range effects on human gene regulation. Elife 2018, 7, e35989.
  144. Steele, A.J.; Al-Chalabi, A.; Ferrante, K.; Cudkowicz, M.E.; Brown, R.H.; Garson, J.A. Detection of serum reverse transcriptase activity in patients with ALS and unaffected blood relatives. Neurology 2005, 64, 454–458.
  145. McCormick, A.L.; Brown, R.H.; Cudkowicz, M.E.; Al-Chalabi, A.; Garson, J.A. Quantification of reverse transcriptase in ALS and elimination of a novel retroviral candidate. Neurology 2008, 70, 278–283.
  146. Li, T.D.; Murano, K.; Kitano, T.; Guo, Y.; Negishi, L.; Siomi, H. TDP-43 safeguards the embryo genome from L1 retrotransposition. Sci. Adv. 2022, 8, eabq3806.
  147. Della Valle, F.; Reddy, P.; Yamamoto, M.; Liu, P.; Saera-Vila, A.; Bensaddek, D.; Zhang, H.; Martinez, J.P.; Abassi, L.; Celii, M.; et al. LINE-1 RNA causes heterochromatin erosion and is a target for amelioration of senescent phenotypes in progeroid syndromes. Sci. Transl. Med. 2022, 14, eabl6057.
  148. Krug, L.; Chatterjee, N.; Borges-Monroy, R.; Hearn, S.; Liao, W.-W.; Morrill, K.; Prazak, L.; Rozhkov, N.; Theodorou, D.; Hammell, M.; et al. Retrotransposon activation contributes to neurodegeneration in a Drosophila TDP-43 model of ALS. PLoS Genet. 2017, 13, e1006635.
  149. Chang, Y.-H.; Dubnau, J. The Gypsy Endogenous Retrovirus Drives Non-Cell-Autonomous Propagation in a Drosophila TDP-43 Model of Neurodegeneration. Curr. Biol. 2019, 29, 3135–3152.e4.
  150. Romano, G.; Klima, R.; Feiguin, F. TDP-43 prevents retrotransposon activation in the Drosophila motor system through regulation of Dicer-2 activity. BMC Biol. 2020, 18, 82.
  151. Kim, V.N.; Han, J.; Siomi, M.C. Biogenesis of small RNAs in animals. Nat. Rev. Mol. Cell Biol. 2009, 10, 126–139.
  152. Pham, J.; Keon, M.; Brennan, S.; Saksena, N. Connecting RNA-Modifying Similarities of TDP-43, FUS, and SOD1 with MicroRNA Dysregulation Amidst A Renewed Network Perspective of Amyotrophic Lateral Sclerosis Proteinopathy. Int. J. Mol. Sci. 2020, 21, 3464.
  153. Aravin, A.A.; Sachidanandam, R.; Bourc’His, D.; Schaefer, C.; Pezic, D.; Toth, K.F.; Bestor, T.; Hannon, G.J. A piRNA Pathway Primed by Individual Transposons Is Linked to De Novo DNA Methylation in Mice. Mol. Cell 2008, 31, 785–799.
  154. Chen, L.; Dahlstrom, J.E.; Lee, S.-H.; Rangasamy, D. Naturally occurring endo-siRNA silences LINE-1 retrotransposons in human cells through DNA methylation. Epigenetics 2012, 7, 758–771.
  155. Gonzalez-Hernandez, M.J.; Swanson, M.D.; Contreras-Galindo, R.; Cookinham, S.; King, S.R.; Noel, R.J.; Kaplan, M.H.; Markovitz, D.M. Expression of Human Endogenous Retrovirus Type K (HML-2) Is Activated by the Tat Protein of HIV-1. J. Virol. 2012, 86, 7790–7805.
  156. Wang, I.-F.; Reddy, N.M.; Shen, C.-K.J. Higher order arrangement of the eukaryotic nuclear bodies. Proc. Natl. Acad. Sci. USA 2002, 99, 13583–13588.
  157. Wang, H.-Y.; Wang, I.-F.; Bose, J.; Shen, C.-K. Structural diversity and functional implications of the eukaryotic TDP gene family. Genomics 2004, 83, 130–139.
  158. D’Alton, S.; Altshuler, M.; Lewis, J. Studies of alternative isoforms provide insight into TDP-43 autoregulation and pathogenesis. RNA 2015, 21, 1419–1432.
  159. Weskamp, K.; Tank, E.M.; Miguez, R.; McBride, J.P.; Gómez, N.B.; White, M.; Lin, Z.; Gonzalez, C.M.; Serio, A.; Sreedharan, J.; et al. Shortened TDP43 isoforms upregulated by neuronal hyperactivity drive TDP43 pathology in ALS. J. Clin. Investig. 2020, 130, 1139–1155.
  160. Shenouda, M.; Xiao, S.; MacNair, L.; Lau, A.; Robertson, J. A C-Terminally Truncated TDP-43 Splice Isoform Exhibits Neuronal Specific Cytoplasmic Aggregation and Contributes to TDP-43 Pathology in ALS. Front. Neurosci. 2022, 16, 868556.
  161. Goossens, J.; Vanmechelen, E.; Trojanowski, J.Q.; Lee, V.M.; Van Broeckhoven, C.; van der Zee, J.; Engelborghs, S. TDP-43 as a possible biomarker for frontotemporal lobar degeneration: A systematic review of existing antibodies. Acta Neuropathol. Commun. 2015, 3, 15.
  162. Yamashita, T.; Akamatsu, M.; Kwak, S. Altered Intracellular Milieu of ADAR2-Deficient Motor Neurons in Amyotrophic Lateral Sclerosis. Genes 2017, 8, 60.
  163. Yang, Z.; Lin, F.; Robertson, C.S.; Wang, K.K. Dual Vulnerability of TDP-43 to Calpain and Caspase-3 Proteolysis after Neurotoxic Conditions and Traumatic Brain Injury. J. Cereb. Blood Flow Metab. 2014, 34, 1444–1452.
  164. Liu, W.; Wang, Z.; Liu, L.; Yang, Z.; Liu, S.; Ma, Z.; Liu, Y.; Ma, Y.; Zhang, L.; Zhang, X.; et al. LncRNA Malat1 inhibition of TDP43 cleavage suppresses IRF3-initiated antiviral innate immunity. Proc. Natl. Acad. Sci. USA 2020, 117, 23695–23706.
  165. Buratti, E. Functional Significance of TDP-43 Mutations in Disease. Adv. Genet. 2015, 91, 1–53.
  166. Mitra, J.; Hegde, M.L. The Role of TDP-43 in Genome Repair and beyond in Amyotrophic Lateral Sclerosis. In Amyotrophic Lateral Sclerosis—Recent Advances and Therapeutic Challenges; Hegde, M.L., Ed.; IntechOpen: London, UK, 2020; ISBN 978-1-83880-580-7.
  167. Fratta, P.; Sivakumar, P.; Humphrey, J.; Lo, K.; Ricketts, T.; Oliveira, H.; Brito-Armas, J.M.; Kalmar, B.; Ule, A.; Yu, Y.; et al. Mice with endogenous TDP -43 mutations exhibit gain of splicing function and characteristics of amyotrophic lateral sclerosis. EMBO J. 2018, 37, e98684.
  168. White, M.A.; Kim, E.; Duffy, A.; Adalbert, R.; Phillips, B.U.; Peters, O.M.; Stephenson, J.; Yang, S.; Massenzio, F.; Lin, Z.; et al. TDP-43 gains function due to perturbed autoregulation in a Tardbp knock-in mouse model of ALS-FTD. Nat. Neurosci. 2018, 21, 552–563.
  169. Arnold, E.S.; Ling, S.-C.; Huelga, S.C.; Lagier-Tourenne, C.; Polymenidou, M.; Ditsworth, D.; Kordasiewicz, H.B.; McAlonis-Downes, M.; Platoshyn, O.; Parone, P.A.; et al. ALS-linked TDP-43 mutations produce aberrant RNA splicing and adult-onset motor neuron disease without aggregation or loss of nuclear TDP-43. Proc. Natl. Acad. Sci. USA 2013, 110, E736–E745.
  170. Ozyurt, T.; Gautam, M. Differential Epigenetic Signature of Corticospinal Motor Neurons in ALS. Brain Sci. 2021, 11, 754.
  171. Hasegawa-Ogawa, M.; Okano, H.J. Characterization of the upstream and intron promoters of the gene encoding TAR DNA-binding protein. Sci. Rep. 2021, 11, 8720.
More
Video Production Service